Surface enhanced Raman spectroscopy on a flat graphene surface Please share

advertisement
Surface enhanced Raman spectroscopy on a flat
graphene surface
The MIT Faculty has made this article openly available. Please share
how this access benefits you. Your story matters.
Citation
Xu, W. et al. “Surface Enhanced Raman Spectroscopy on a Flat
Graphene Surface.” Proceedings of the National Academy of
Sciences 109.24 (2012): 9281–9286. Web.© 2012 National
Academy of Sciences.
As Published
http://dx.doi.org/10.1073/pnas.1205478109
Publisher
National Academy of Sciences (U.S.)
Version
Final published version
Accessed
Wed May 25 22:02:31 EDT 2016
Citable Link
http://hdl.handle.net/1721.1/76754
Terms of Use
Article is made available in accordance with the publisher's policy
and may be subject to US copyright law. Please refer to the
publisher's site for terms of use.
Detailed Terms
Surface enhanced Raman spectroscopy
on a flat graphene surface
Weigao Xua, Xi Linga, Jiaqi Xiaoa, Mildred S. Dresselhausb,c, Jing Kongb, Hongxing Xud, Zhongfan Liua, and Jin Zhanga,1
a
Center for Nanochemistry, Beijing National Laboratory for Molecular Sciences, Key Laboratory for the Physics and Chemistry of Nanodevices, State Key
Laboratory for Structural Chemistry of Unstable and Stable Species, College of Chemistry and Molecular Engineering, Peking University, Beijing 100871,
China; bDepartment of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139; dBeijing National
Laboratory of Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China; and cDepartment of Physics,
Massachusetts Institute of Technology, Cambridge, MA 02139
Surface enhanced Raman spectroscopy (SERS) is an attractive
analytical technique, which enables single-molecule sensitive detection and provides its special chemical fingerprints. During the
past decades, researchers have made great efforts towards an ideal
SERS substrate, mainly including pioneering works on the preparation of uniform metal nanostructure arrays by various nanoassembly and nanotailoring methods, which give better uniformity and
reproducibility. Recently, nanoparticles coated with an inert shell
were used to make the enhanced Raman signals cleaner. By depositing SERS-active metal nanoislands on an atomically flat graphene
layer, here we designed a new kind of SERS substrate referred to as
a graphene-mediated SERS (G-SERS) substrate. In the graphene/
metal combined structure, the electromagnetic “hot” spots (which
is the origin of a huge SERS enhancement) created by the gapped
metal nanoislands through the localized surface plasmon resonance
effect are supposed to pass through the monolayer graphene, resulting in an atomically flat hot surface for Raman enhancement.
Signals from a G-SERS substrate were also demonstrated to have
interesting advantages over normal SERS, in terms of cleaner vibrational information free from various metal-molecule interactions
and being more stable against photo-induced damage, but with
a comparable enhancement factor. Furthermore, we demonstrate
the use of a freestanding, transparent and flexible “G-SERS tape”
(consisting of a polymer-layer-supported monolayer graphene with
sandwiched metal nanoislands) to enable direct, real time and reliable detection of trace amounts of analytes in various systems,
which imparts high efficiency and universality of analyses with
G-SERS substrates.
atomically smooth substrate ∣ metal-molecule isolation ∣ signal fluctuation ∣
mediator ∣ application
F
or almost all sorts of analytical methods, the dream to improve
their sensitivity as well as their reproducibility and to optimize
the analytical process (e.g., to simplify the sample preparation/
measurement procedures for quick analysis and to enable in situ
and real time monitoring) is a long-term pursuit. Spectroscopic
approaches based on fluorescence, infrared absorption and
Raman scattering have been developed with rising importance
for various sensing and imaging applications. Among them,
Raman scattering provides more structural information (characteristic vibrational information of each chemical bond) over
fluorescence, and higher spatial resolution (shorter excitation
wavelength) over infrared absorption (1). However, as Raman
scattering is an inelastic scattering process with a very low crosssection, it is not very sensitive and thus has limited analysis efficiency and applicability.
Because of this, surface enhanced Raman spectroscopy
(SERS) has been developed since the 1970s (2–4) to enable ultrasensitive characterization down to the single-molecule level (5,
6), comparable to single-molecule fluorescence spectroscopy (7).
In a SERS experiment, a rough metal surface or colloidal metal
nanostructures are typically used. Here, the highly curved or
gapped metal regions (under proper incident light conditions)
www.pnas.org/cgi/doi/10.1073/pnas.1205478109
give rise to greatly enhanced local electromagnetic fields via
the localized surface plasmon resonance effect. The localized
electromagnetic field (usually called a “hot” spot) thus results in
the dominant contribution of SERS enhancement [referred to as
electromagnetic enhancement (EM)], which can reach 10 8 or
more (8–10). Usually there is another chemical contribution
called chemical enhancement (CM), but with a minor enhancement factor (typically 10 to 10 2 ) (11, 12). As a result, molecules
adsorbed near the electromagnetic hot spots dominate the Raman intensity. The distribution of molecules on a normal SERS
substrate is usually complicated, as the molecules near the hot
spots can be in fluctuating amounts and random orientations.
On the other hand, chemical interactions between the molecules
and the metal substrate can make this case more complicated.
Chemical adsorption-induced vibrations, molecular deformation
and distortion, charge transfer between the metal and molecules,
photocarbonization, photobleaching or metal-catalyzed side reactions (10, 13) may all affect the final signal and make it difficult
to precisely assign each vibration of a SERS spectrum. Thus,
the SERS signal is sometimes too sensitive to a SERS substrate,
possible fluctuations caused by nonuniform molecular adsorption
(amounts/configurations), different metal enhancers and the
corresponding metal-molecule interactions bring down the comparability between parallel SERS experiments, and as a result
standard analytical methods based on SERS have been rarely
popularized. To further understand the SERS effect and fully
excite its advanced applications, a SERS substrate with a more
“intelligent” form is still lacking.
From the above, the availability of an enhancement substrate,
which can give a uniform, stable, clean and highly sensitive SERS
response has remained a bottleneck for extending the further
applications of SERS (10, 14). Current efforts towards more
reproducible SERS analyses are mainly focused on the nanotailoring or nanoassembly of substrates with uniform metal nanostructure arrays (15–19). However, a “uniform” SERS substrate is
still not uniform on the nanoscale, the distribution of molecules
and their states may be quite uncontrollable in the vicinity of the
electromagnetic hot spots. On the other hand, metal nanoparticles with a thin and inert SiO2 shell were recently demonstrated
to make SERS signals cleaner (13). Yet, the coating of a pinholefree and thin SiO2 layer (to avoid metal-molecule interaction
without large sacrifice to the SERS-activity) is an over delicate
technique. In this work, based on the use of the two-dimensional
atomic crystal graphene, we designed a unique kind of SERS
Author contributions: W.X., X.L., and J.Z. designed research; W.X., X.L., J.X., and H.X.
performed research; H.X. and J.Z. contributed new reagents/analytic tools; W.X., X.L., J.X.,
M.S.D., J.K., Z.L., and J.Z. analyzed data; and W.X., X.L., J.X., M.S.D., J.K., and J.Z. wrote the
paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
1
To whom correspondence should be addressed. E-mail: jinzhang@pku.edu.cn.
This article contains supporting information online at www.pnas.org/lookup/suppl/
doi:10.1073/pnas.1205478109/-/DCSupplemental.
PNAS ∣ June 12, 2012 ∣
vol. 109 ∣
no. 24 ∣
9281–9286
CHEMISTRY
Edited by Nicholas J. Turro, Columbia University, New York, NY, and approved April 18, 2012 (received for review April 3, 2012)
D
monolayer graphene
(1LG)
metal
nanoislands
Normal SERS
A
supporting layer
strong
C
1LG
E
0.5 nm
Electromagnetic field
B
1LG
Graphene-mediated
SERS (G-SERS)
G-SERS tape
1LG
Au
Au/Ag
weak
Fig. 1. Design of a G-SERS substrate. (A) Components of a G-SERS tape. (B) The simulated electromagnetic field distribution of a 60-nm gold hemisphere dimer
covered with graphene (with a 2-nm gap between two gold hemispheres and a 0.5-nm gap between gold hemispheres and graphene). The excitation
wavelength is 632.8 nm, with the incident direction perpendicular to the graphene surface (from top to bottom). (C) The resulting “hot” graphene surface
to serve as a G-SERS substrate. (D, E) Schematic illustration of molecules adsorbed on a normal SERS substrate (D) and on a G-SERS substrate (E).
substrate containing a graphene/metal-combined structure, referred to as a graphene-mediated SERS (G-SERS) substrate.
Results
As illustrated in Fig. 1, the active surface of our G-SERS
substrate is a monolayer graphene (1LG), with gold/silver nanoislands tightly adhered on the backside. By collecting electromagnetic hot spots from the metal nanoislands on a graphene surface,
a G-SERS substrate is anticipated to offer an atomically smooth
surface for controllable molecular arrangements as well as welldefined molecular interactions, yet also has the electromagnetic
hot spots created by the localized surface plasmon resonance of
gold or silver nanoislands adhered on the backside of graphene.
In a G-SERS substrate, the metal nanoislands focus the incident laser, creating localized electromagnetic hot spots close to
the graphene surface. The permeability of the electromagnetic
field to pass through the graphene is critical for the final enhancement activity. First, theoretical calculations are carried out. For
simplicity, two 60-nm diameter gold hemispheres adhering below
a monolayer graphene are used to simulate a G-SERS substrate.
From the three dimensional finite difference time domain (3DFDTD) simulations results as shown in Fig. 1B, localized electromagnetic field occurs at the gap of the hemisphere dimer, spreads
from the central strongest hot spot to the vicinity (The permeability was also supported by extended Mie theory calculations of a
graphene-coated 60-nm gold sphere dimer, SI Appendix, Fig. S1).
By comparing with that of gold hemispheres without graphene,
we find that graphene exhibits good permeability for the electromagnetic hot spot to pass through, and interestingly the maximum jEloc ∕E0 j 2 is even almost one time stronger (SI Appendix,
Fig. S2). These results encouraged us that the close vicinity of
these hot spots from metal nanostructures can make the monolayer graphene a hot surface (Fig. 1C), and thus allows localized
electromagnetic enhancement to be achieved on an atomically
flat surface.
Fig. 1D and 1E illustrate possible states of molecules adsorbed
on a normal SERS substrate (Fig. 1D) and a G-SERS substrate
(Fig. 1E). In Fig. 1D, molecules are in various orientations
according to the morphology of the metal nanoislands. The enlarged view in Fig. 1D illustrates various possibilities that may
make the enhanced Raman signals complicated. In contrast, by
taking the flat and hot graphene surface as the Raman enhancer,
as illustrated in Fig. 1E, the strong Raman enhancement from
localized electromagnetic hot spots created by metal nanoislands
can be realized on an atomically smooth surface. The monolayer
graphene is expected to act as a multifunctional “SERS mediator”, including: (i) a flat supporting surface to arrange molecules
9282 ∣
www.pnas.org/cgi/doi/10.1073/pnas.1205478109
in a more controllable way, (ii) a spacer to separate the metalmolecule contact, (iii) additional effects like a stabilizer of both
the substrate and the molecules under laser exposure.
To seek the possible superiorities of a G-SERS substrate, a
rational way was used to fabricate substrates with both G-SERS
regions and normal SERS regions for comparison. As shown in
Fig. 2A, a uniform layer of molecules, such as rhodamine 6G
(R6G) or copper phthalocyanine (CuPc), were first deposited
on a SiO2 ∕Si substrate by vacuum thermal evaporation. Mechanically exfoliated graphene pieces were then transferred to the top
of the molecules, creating regions with and without graphene (for
regions without graphene, our Raman inspections were carried
out only on those regions that are not touched by the adhesive
scotch tape used in the mechanical exfoliation, to make sure
no damage was done to the Raman probes). Finally, gold or silver
nanoislands adhering to the graphene were constructed by
vacuum thermal evaporation of an 8-nm metal film (a preferred
thickness for gold as illustrated in SI Appendix, Figs. S3 and S4)
on the top of both regions (1LG and the bare substrate). It should
be pointed out that gold and silver are the most popular electromagnetic enhancers, where gold is more stable but the enhancement is usually less than that for silver. Silver is more likely to be
affected by the environment, such as oxidation processes which
will decrease the enhancement over time. Here we show that both
of gold and silver can be used effectively for the fabrication of a
G-SERS substrate. The method of vacuum thermal evaporation
can avoid external contaminations such as surfactants which
would be introduced by using nanostructures prepared by colloidal methods (20). Thus, G-SERS and normal SERS regions
(defined as G-SERS(M) and SERS(M), where M is the metal
used as the electromagnetic enhancer) were constructed corresponding to the regions with and without graphene, respectively.
Typical Raman spectra of R6G with G-SERS(Au), G-SERS
(Ag), SERS(Au), and SERS(Ag) are shown in Fig. 2B from bottom to top, respectively. First, it can be seen that the G-SERS
regions give more reproducible and cleaner signals as compared
to normal SERS. For G-SERS(Au) and G-SERS(Ag), the spectral features are the same as one another, with clear peaks at 612,
635, 660, 771, 1181, 1313, 1361, 1511, 1648 cm −1 (1 cm −1 ) (see
SI Appendix, Table S1). The peaks at around 1570 and 2625 cm −1
(marked by *) are the G-band and G′-band of monolayer graphene, respectively. The relative intensity of the graphene peak
as compared to R6G is different in G-SERS(Au) and G-SERS
(Ag) (this variability is reproducible from sample to sample in our
parallel measurements); this may be because of a different chemical contribution of gold and silver. In contrast, Raman spectra
enhanced by SERS(Au) and SERS(Ag) show many additional
Xu et al.
A
B
N
N
Cu
CuPc
SiO2/Si
graphene
transferred
Au/R6G
SERS (Au) 2.5 mW
Ag/1LG/R6G
G-SERS (Ag) 0.25 mW
*
SiO2/Si
G-SERS
SERS
Raman intensity (a.u.)
C
metal film
deposited
*
G-SERS (Au) 2.5 mW
750
SiO2/Si
1,000
SERS-active metal film
1,250
1,500
-1
Raman shift (cm )
SERS
1,451
Au/R6G
Au/1LG/R6G
*
1,600
2,400
-1
Raman shift (cm )
953
G-SERS
1,453
953
2,750
SERS 953
SERS 1,451
G-SERS 953
G-SERS 1,453
0
200 400
Time (s)
960 1,440 1,560
-1
Raman shift (cm )
peaks (as marked by red arrows) compared to the G-SERS results. Among these additional peaks, such as at 597, 1014,
1411, 1468, 1540 cm −1 in SERS(Au) and 554, 590, 1012, 1336,
1403, 1471, 1531 cm −1 in SERS(Ag), some of them may be from
contamination of surface carbons (as will be described below),
some are new emerging bands possibly due to the modified surface selection rules (because of alteration of the symmetry of the
vibrational modes under changed configurations), while others
remain difficult to assign precisely. Parallel measurements are
demonstrated and we found that these additional signals are irreproducible (see SI Appendix, Fig. S6). When looking at those
bands, which are also observed in the G-SERS spectra, many
bands are shifted (probably due to a charge-transfer effect of the
metal-R6G), and the amount of shifts are different for SERS(Au)
and SERS(Ag). Thus, the SERS results will depend on various
factors, which may result in the appearance and disappearance
of various features, mixed with peaks having up or down shifted
frequencies. As a result, it would be almost impossible to assign
each vibration precisely, since the spectral features are no longer
clearly dependent on the primary structure of the sample molecules.
Additionally, we also found that graphene makes the SERS
system more stable. Photo-induced damage is a common characteristic of SERS, and is popularly known as photocarbonization
and photobleaching. Raman spectra of the as prepared G-SERS
and SERS substrates (without molecules) are shown in Fig. 2C,
top panel. First, one can see that the G-SERS substrate provides
a cleaner baseline than that of normal SERS: The photocarbonization background in the 1100 ∼ 1700 cm −1 range (marked
in blue line) for the SERS spectra is not observed in the G-SERS
spectra, together with the emergence of a characteristic signal of
monolayer graphene. Because no Raman probes were introduced
here, this carbonization background is normally considered to be
due to surface carbon-based adsorbates from the atmosphere. In
fact, the photocarbonization of adsorbates was observed and studied soon after the SERS effect was discovered (21), but so far
this signal contamination problem has not yet been resolved. This
photocarbonization effect will be even more pronounced in an
Xu et al.
2,500
*
*
800
*
D
Au
Au/1LG
*
Raman probes
Monolayer graphene (1LG)
*
Au/1LG/R6G
600
Fig. 2. Comparison of signals from a G-SERS substrate
and a normal SERS substrate. (A) Schematic route for
sample preparation with G-SERS and SERS regions on
a SiO2 ∕Si substrate. An enlarged view shows the chemical structure of the probe molecules used in the
experiment. (B) SERS and G-SERS spectra of R6G with
gold and silver nanoislands used as an electromagnetic enhancer, respectively. Red arrows here point to
additional and nonreproducible peaks in spectra of
normal SERS. (C) Photocarbonization effect (which
causes a background at 1100 ∼ 1700 cm −1 ) in SERS
and G-SERS substrates. (D) Stability of the SERS (top)
and G-SERS (bottom) spectra of CuPc in a time series
of 600 s. The inset shows the integrated intensity
of the peaks (at 953;1451 cm −1 for SERS, and
953;1453 cm −1 for G-SERS, respectively) change with
the increased acquisition time. Time sequence: from
top to bottom. “*” in B and C marks the G-band
(∼1570 cm −1 ) and G′-band (∼2625 cm −1 ) of 1LG.
G-SERS spectra are measured under the same conditions with the corresponding SERS spectra and are
shown on the same intensity scale, except for C (bottom), where the SERS spectrum was acquired under a
10% laser power compared to the G-SERS spectrum.
actual SERS measurement when probe molecules are present.
As shown in Fig. 2C (bottom), even with a much lower laser
power (the SERS spectrum is acquired under a 10% laser power
compared to G-SERS), the SERS spectrum of R6G shows an
obvious photocarbonization background (marked in blue line)
while this background is absent in the G-SERS spectrum. The
presence of graphene cover layer may suppress the catalytic activity of the gold nanoislands and prevent the photo-induced carbonization, as well as keep R6G molecules from direct contact with
the gold film, thus preventing adsorbates (from the atmosphere)
and R6G molecules from photocarbonization. The high thermal
conductivity of graphene may also partly contribute to the enhanced stability, and actually we have found that the morphology
of the metal film changes after a relatively large laser power
exposure if there is no graphene layer present, while it keeps
stable for the G-SERS regions. An additional effect, the lower
photoluminescence background in a G-SERS substrate than that
of normal SERS, is consistent with the reported results that graphene is an effective quencher for both the fluorescence of dyes
(22, 23) and the photoluminescence of a gold film, through a resonance energy transfer process (24).
On the other hand, the photobleaching (or photodegradation)
of the Raman probes induced by the laser is also a well-known
feature in normal SERS experiments, especially for dyes. During
a normal SERS measurement, the intensity of the Raman signals
of molecules may decrease with increased acquisition time (especially under a high laser power), and this effect will lead to a variation of the Raman spectra with acquisition time and laser
power. Fig. 2D shows a comparison of the stability of the Raman
spectra of CuPc of SERS(Au) (top) and G-SERS(Au) (bottom)
regions. It is seen here that the signal of CuPc in the normal
SERS region decreases quickly in intensity during a 600 s measurement while in the G-SERS region it remains totally stable. It
is speculated that both the separation of CuPc from the rough
gold film and the formation of a graphene/CuPc complex through
π-π interactions may contribute to the stability of CuPc.
The above results in Fig. 2 suggest that, in a G-SERS substrate
(when the rough metal substrate is replaced by a hot and atomPNAS ∣
June 12, 2012 ∣
vol. 109 ∣
no. 24 ∣
9283
CHEMISTRY
R6G
N
N
Peak area (a.u.)
N
H
O
N
time
N
time
N
Raman intensity (a.u.)
N
H
or
Raman intensity (a.u.)
O
OCI
Ag/R6G
SERS (Ag) 0.25 mW
N
ically flat graphene surface), Raman enhancement turns out to be
cleaner (with characteristic and reproducible vibrational modes
in which fluctuating information is removed) and more stable
(against laser-induced damage such as photocarbonization and
photobleaching). It should be emphasized that the above characteristics of a normal SERS substrate do not mean shortcomings of
the SERS technique itself, and actually it oppositely reflects the
fascinating capability of SERS (even capable of monitoring the
actions of a single molecule). However, currently the state of
the art of SERS is still mostly limited to the study of a mixture
of molecules in unclear states with poor controllability. Thus, the
G-SERS substrate is anticipated to act as an effective tool to implement SERS, both for general analyses (with better controllability and reproducibility among parallel measurements) and
for many advanced applications.
Further, we checked the enhancement factor of G-SERS
substrates. For comparison, the Raman spectra of CuPc (pristine), 1LG/CuPc [graphene-enhanced Raman scattering (25),
“GERS”], Au/CuPc [SERS(Au)], Au/1LG/CuPc [G-SERS(Au)],
Ag/CuPc [SERS(Ag)], and Ag/1LG/CuPc [G-SERS(Ag)] are
shown in Fig. 3(A–F). All the spectra are acquired under the same
conditions and all the spectra are shown with the same intensity
scale, where the Raman signal data is multiplied by 30 in the case
of the pristine and GERS and by 10 in the case of G-SERS(Au)
and SERS (Au). Here, the relative intensity of the graphene signal seems much lower (the G-band of graphene is almost invisible
but the G′-band can still be clearly distinguished) as compared to
measurement of R6G in Fig. 2B, and this is because of a resonance Raman enhancement effect of CuPc under our measurement condition (632.8 nm). Through the integrated area of the
SERS(Ag)
F
Ag
x 10
Ag
x 10*
G-SERS(Ag)
Raman intensity (a.u.)
E
D
SERS(Au)
x 10
C
G-SERS(Au)
x 10
x 100
Au
x 100
*
Au
GERS
B
x 30
A
pristine
x 30
x 300
CuPc
600
900
x 300
1,200 1,500 1,800 2,100 2,400 2,700
-1
Raman shift (cm )
Fig. 3. Comparison of the Raman spectra of CuPc with different enhancement methods. (A) Pristine spectrum: CuPc∕SiO2 ∕Si. (B) GERS spectrum: 1LG∕
CuPc∕SiO2 ∕Si. (C, D) G-SERS (Au∕1LG∕CuPc∕SiO2 ∕Si) (C) and SERS (Au∕CuPc∕
SiO2 ∕Si) (D) spectra of CuPc with Au as the electromagnetic enhancer. (E, F)
G-SERS (Ag∕1LG∕CuPc∕SiO2 ∕Si) (E) and SERS (Ag∕CuPc∕SiO2 ∕Si) (F) spectra of
CuPc with Ag as the electromagnetic enhancer. In A and B the spectra were
magnified by a factor of 30. In C and D the spectra were magnified by a factor
of 10. Inset pictures are the corresponding sample structures that represent
different enhancement methods. An enlarged view of the regions at 2400 ∼
2720 cm −1 (including the G′-band of 1LG) is also shown, with G′-band of 1LG
marked by “*”.
9284 ∣
www.pnas.org/cgi/doi/10.1073/pnas.1205478109
1530 cm −1 peak, taking the pristine spectra as a reference, it is
enhanced by a factor of 14 for GERS, 61 for SERS(Au), 85 for
G-SERS(Au), 580 for SERS(Ag), and 755 for G-SERS(Ag). It
should be noted that here the enhancement factors for the SERS
region and the G-SERS region are both seriously undervalued
since the absorption of the metal film makes only molecules at the
gaps between the metal nanoislands (of course, this is quite a
minority) contribute to their final signal intensity. However, here
it is fair to compare the relative sensitivity between a normal
SERS substrate and a G-SERS substrate. The weak signal of pristine CuPc is magnified, and under the measurement conditions
(632.8 nm excitation), silver is a better electromagnetic enhancer
than gold. Most importantly, for both the case of gold and silver,
the enhancement factors of the G-SERS regions are even slightly
higher than those of normal SERS for most of the bands. This
means that the presence of the graphene layer will not cause the
localized electromagnetic field to decay observably in its vicinity,
which is consistent with our theoretical results as mentioned
before. The small enhancement of the graphene substrate is consistent with our earlier results (25, 26), possibly due to a chemical
enhancement. By a careful comparison of their spectral features,
it can be seen that the spectral feature from a G-SERS substrate
remains the same as that from a graphene substrate while the
enhancement was greatly improved by the introduction of the
electromagnetic enhancement. Consistent results can be observed from the signals of R6G from a G-SERS substrate and
a graphene substrate as shown in Table S1 (SI Appendix), regardless of the material of the metal enhancer. Thus, by collecting hot
spots of metal nanoislands on a flat graphene surface, G-SERS
substrate provides comparable sensitivity with that of normal
SERS. The better-defined molecule-substrate interactions in a
G-SERS substrate is also intriguing, as an atomically compact
and chemically inert graphene surface seems to cut down signal
fluctuations among different metal enhancers.
A “universal” substrate should be sufficiently economical,
sufficiently convenient, and sufficiently compatible with various
samples in a different state. As described in the following, we
developed a general type of G-SERS substrate (which we call a
“G-SERS tape”), which is transparent, freestanding, and flexible
(by virtue of the high light transmittance and good flexibility
of both graphene and the selected polymer). As a schematic
representation in Fig. 1A, a G-SERS tape consists of three parts:
monolayer graphene, a polymer supporting layer, and sandwiched metal nanoislands adhering to the backside of the active
face of graphene. Fabrication of G-SERS tapes and their applications in various situations, including surfaces with arbitrary
morphology, or even directly on a solution surface are demonstrated. Such a G-SERS process does not require special sample
preparation, is a noninvasive process, and is very convenient for
in situ characterization or real time monitoring.
As shown in Fig. 4A, taking the G-SERS(Au) tape as example,
an 8-nm gold film was first deposited on a large area monolayer
graphene grown on a 8 × 8 cm 2 copper foil [prepared by the chemical vapor deposition method (27)], which is uniform with high
quality (see SI Appendix, Fig. S7). A poly(methyl methacrylate)
(PMMA) film, which is optically transparent, inert, and Raman
inactive, was then fabricated by spin-coating and dried at 170 °C
for 30 min (27, 28). Finally, a transparent, freestanding, and flexible G-SERS tape with a PMMA/metal/graphene structure was
obtained by etching the copper foil in a FeCl3 ∕H2 O solution. This
technique is suitable for scaled production of G-SERS tapes
(Fig. 4C). The transmission electron microscopy (TEM) image in
Fig. 4B shows the morphology of an 8-nm gold film, in which most
gaps between the nanoislands are 2 ∼ 3 nm, thereby allowing for a
considerable electromagnetic enhancement. As shown in Fig. 4D,
the atomic force microscopy characterization shows that the active
side (graphene side) of the G-SERS tape is flat, with a fluctuation
(surface roughness) along the film length of less than 2 nm.
Xu et al.
A graphene
B
C
adding a PMMA layer
by spin coating
removal of the
copper in Fe3+/H20
D
Fig. 4. Transparent, freestanding and flexible
G-SERS tape prepared from CVD-grown monolayer
graphene. (A) Schematic steps of the preparation
route. (B) TEM images of an 8-nm gold film with a
global view (top) and an enlarged view (bottom),
showing that most gaps are on the size of 2 ∼ 3 nm.
(C) Photograph of an 8 × 8 cm 2 G-SERS substrate
before the removal of copper. Insets on the right are
two freestanding G-SERS(Au) tapes, one is floating
on water (top) and the other is held with tweezers
(supported by a windowed scotch tape) (bottom), respectively. (D) Atomic force microscopy image of the
graphene side of a G-SERS(Au) tape (top), together
with a section analysis on the bottom. This indicates
that the surface of the resulting G-SERS tape is flat,
with a height fluctuation of less than 2 nm.
10.0 nm
50 nm
5 nm
Height (nm)
2 mx 2 m
2
0
-2
0.0
Mechanically exfoliated graphene is also suitable for use in
fabricating G-SERS tapes (see SI Appendix, Fig. S8). A G-SERS
tape with silver as an electromagnetic enhancer [G-SERS(Ag)
tape)] are prepared through a similar procedure (with a few modifications) and this tape is also described in the SI Appendix. As
the active metal is buried inside the graphene and the PMMA
layer (in the case of Ag, there is an additional protection layer)
without exposure to the atmosphere, it is anticipated that these
G-SERS tapes could have a long shelf life. Because of the flexibility of a G-SERS tape, the contact between the tape and
the target material will be good enough to achieve the G-SERS
enhancement.
Fig. 5 shows several typical analyses. As shown in Fig. 5A,
G-SERS tapes shown in Fig. 4 can be used directly for real time
analysis of samples in an aqueous solution, where we first put
such a G-SERS tape floating on water, and a clean baseline was
obtained with a clear graphene signal. Next, the same G-SERS
tape was placed on a 1 × 10 −5 M aqueous solution of R6G, and
the intrinsic signal of R6G appeared in the Raman spectra (with
the same spectral features as the G-SERS results in Fig. 2B). Additionally, the R6G signal disappeared after washing and replacing the G-SERS tape on water. This reversibility suggests that
G-SERS tape may be exploited in real time sensing processes,
such as online monitoring of water contaminants. Furthermore,
the Raman spectra of a self-assembled monolayer of p-aminothiophenol on a flat gold surface (Fig. 5B) and of CuPc molecules adsorbed on a cauliflower surface (Fig. 5C) shows that
G-SERS tapes can be used for the analysis of trace amounts of
molecules on any surface with arbitrary morphology. Moreover,
we extend G-SERS measurements to other solvents like ethanol
by making a small modification to the G-SERS tape (see SI
Appendix, Fig. S13). Therefore, this kind of G-SERS substrate
is widely applicable for direct, noninvasive and ultrasensitive
Raman measurements, including real time monitoring on a solution surface and on other surfaces with any arbitrary morphology.
Discussion
SERS is a powerful tool and it can be more powerful when
exploited with a carefully designed substrate. Among the wide
researches on SERS during the past years, great efforts have been
made towards gaining a theoretical understanding of the enhancement mechanism (4, 29, 30) and resulting in extended practical applications (13, 31–33). The basic physical characteristic
that a localized surface plasmon resonance (the origin of a large
SERS enhancement) occurs only at those closely gapped or
highly curved plasmonic metal nanostructures has hindered the
SERS technique from being used on a flat surface. In this report,
we demonstrated a G-SERS substrate in which a plasmonic metal
Xu et al.
0.5
1.0
1.5
2.0
nanoisland array is adhered to a graphene surface, creating an
atomically smooth surface with strong electromagnetic hot spots
that can be used for Raman enhancement. The concept of enhancing the Raman signals on a flat graphene surface simplifies the
A
I G-SERS tape on H2O
*
II moved onto R6G/H2O
*
III replaced on H2O
*
CHEMISTRY
deposition of
metal nanoislands
800
1,200
1,600
-1
Raman shift (cm )
B
*
*
NH2NH2NH2
1,074
1,184
copper foil
G-SERS tape
G-SERS
pristine
S
S
x 20
S
Au
1,200
1,600
2,550
-1
Raman shift (cm )
C
G-SERS
*
*
pristine
800
1,600
2,400
-1
Raman shift (cm )
Fig. 5. Analyzing trace amounts of molecules under a variety of situations
using the transparent, freestanding and flexible G-SERS(Au) substrate. (A) A
real time and reversible G-SERS characterization of R6G directly in a
1 × 10 −5 M aqueous solution. (I, II, III are the Raman spectra with the same
G-SERS tape on H2 O, R6G∕H2 O and replaced on H2 O, respectively). (B) Pristine
(red line) and G-SERS (black line) measurements of a self-assembled monolayer of p-aminothiophenol on a flat gold surface. (C) Pristine (red line) and
G-SERS (black line) measurements of a cauliflower surface with adsorbed
CuPc (by soaking in a 1 × 10 −5 M CuPc solution in ethanol for 10 min).
“*” marks the enhanced G-band and G’-band features of 1LG. Sample structures are shown in the left insets.
PNAS ∣
June 12, 2012 ∣
vol. 109 ∣
no. 24 ∣
9285
determination of the precise amount of molecules, their detailed
position and orientation information, as well as the moleculesubstrate interaction. The characteristic Raman signal and atomically precise amount of graphene also make it a natural internal
standard. G-SERS substrates are anticipated to be exploited in
many fields, such as reliable qualitative and label-free quantitative detection, and the characterization of intrinsic structural
information from samples without the introduction of spurious
spectral features coming from metal contamination.
In conclusion, by introducing plasmonic metal to a graphene
layer, an atomically flat substrate for SERS was developed. Cleaner and more reproducible signals were obtained with a comparable enhancement factor. The use of a transparent, freestanding
and flexible G-SERS tape also opens up opportunities for more
challenging applications such as direct, noninvasive and online
analyses for samples in various states. G-SERS substrate shows
a unique application of graphene and it is anticipated to benefit
both the deeper understanding of the SERS effect itself and various practical applications of SERS.
(34). Monolayer graphenes (1LGs) were identified by optical microscopy
(OM) and Raman spectroscopy. Copper phthalocyanine (CuPc) from AlfaAesar, Rhodamine 6G (R6G) from Sigma-Aldrich and Crystal Violet (CV) from
Sinopharm Chemical Reagent Co., Ltd. were used as Raman probes directly
as received. PMMA solid (996 K, from Sigma-Aldrich) was dissolved in ethyl
lactate (5 wt%) for spin coating. Gold and silver films were deposited at a rate
of about 1 Å∕s, and at a pressure of about 10 −3 Pa by vacuum thermal
evaporation. All Raman spectra were obtained using a Horiba HR800 Raman
system with a 632.8 nm He-Ne laser line, diameter of the laser spot is about
1 μm. All spectra in comparison were obtained under the same conditions if
not specially pointed out. TEM characterizations of gold and silver films were
carried out on a TEKNAI F20 equipment, the accelerating voltage is 200 kV.
Samples were prepared by vacuum thermal evaporation of an 8-nm gold or
silver film onto a copper grid (carbon covered). Three-dimensional-FDTD
simulations were carried out by commercial software FDTD solutions 6.5
(Lumerical). The dielectric functions used to describe metal and graphene
are from K. S. Novoselov, et al. (35, 36) for monolayer graphene and are from
P. B. Johnson and R. W. Christy (37) for gold and silver.
Mechanically exfoliated graphene was prepared from Kish graphite (Covalent Materials Corp.) using Scotch tape on a SiO2 ð300 nmÞ∕Si substrate
ACKNOWLEDGMENTS. We thank W.Q. Cai, T. Gao, Y.B. Chen, L. Zhou,
C.H. Zhang, L. Chen, S. P. Zhang for assistance in the experiments. This work
was supported by MOST (2011YQ0301240201 and 2011CB932601), NSFC
(51121091, 50972001, and 21129001), and Fundamental Research Funds for
the Central Universities. M. S. Dresselhaus acknowledges support from NSF/
DMR-1004147. J. Kong acknowledges support from NSF/DMR-0845358.
1. Freudiger CW, et al. (2008) Label-free biomedical imaging with high sensitivity by
stimulated Raman scattering microscopy. Science 322:1857–1861.
2. Fleischmann M, Hendra PJ, McQuillan AJ (1974) Raman spectra of pyridine adsorbed at
a silver electrode. Chem Phys Lett 26:163–166.
3. Albrecht MG, Creighton JA (1977) Anomalously intense Raman spectra of pyridine at a
silver electrode. J Am Chem Soc 99:5215–5217.
4. Jeanmaire DL, Van Duyne RP (1977) Surface Raman spectroelectrochemistry Part I.
Heterocyclic, aromatic, and aliphatic amines adsorbed on anodized silver electrode.
J Electroanal Chem 84:1–20.
5. Kneipp K, et al. (1997) Single molecule detection using surface-enhanced Raman
scattering (SERS). Phys Rev Lett 78:1667–1670.
6. Nie SM, Emory SR (1997) Probing single molecules and single nanoparticles by surfaceenhanced Raman scattering. Science 275:1102–1106.
7. Orrit M, Bernard J (1990) Single pentacene molecules detected by fluorescence excitation in a p-terphenyl crystal. Phys Rev Lett 65:2716–2719.
8. Otto A, Mrozek I, Grabhorn H, Akemann W (1992) Surface-enhanced Raman scattering. J Phys Condens Matter 4:1143–1212.
9. Schatz GC, Young MA, Van Duyne RP (2006) Electromagnetic mechanism of SERS. Top
Appl Phys 103:19–46.
10. Le Ru EC, Etchegoin PG (2009) Principles of Surface-Enhanced Raman Spectroscopy and
related plasmonic effects (Elservier, Amesterdam).
11. Jiang XD, Campion A (1987) Chemical effects in surface-enhanced Raman scattering:
Pyridine chemisorbed on silver adatoms on Rh(100). Chem Phys Lett 140:95–100.
12. Udagawa M, Chou CC, Hemminger JC, Ushioda S (1981) Raman scattering cross section
of adsorbed pyridine molecules on a smooth silver surface. Phys Rev B 23:6843–6846.
13. Li JF, et al. (2010) Shell-isolated nanoparticle-enhanced Raman spectroscopy. Nature
464:392–395.
14. Lin XM, Cui Y, Xu YH, Ren B, Tian ZQ (2009) Surface-enhanced Raman spectroscopy:
substrate-related issues. Anal Bioanal Chem 394:1729–1745.
15. Freeman RG, et al. (1995) Self-assembled metal colloid monolayers: An approach to
SERS substrates. Science 267:1629–1632.
16. Qin LD, et al. (2006) Designing, fabricating, and imaging Raman hot spots. Proc Natl
Acad Sci USA 103:13300–13303.
17. Im H, Bantz KC, Lindquist NC, Haynes CL, Oh SH (2010) Vertically oriented sub-10-nm
plasmonic nanogap arrays. Nano Lett 10:2231–2236.
18. Lim DK, Jeon KS, Kim HM, Nam JM, Suh YD (2010) Nanogap-engineerable Ramanactive nanodumbbells for single-molecule detection. Nat Mater 9:60–67.
19. Lim DK, et al. (2011) Highly uniform and reproducible surface-enhanced Raman scattering from DNA-tailorable nanoparticles with 1-nm interior gap. Nat Nanotechnol
6:452–460.
20. Frens G (1973) Controlled nucleation for regulation of particle size in monodisperse
gold suspensions. Nature Phys Sci 241:20–22.
21. Cooney RP, Mahoney MR, Howard MW (1980) Intense Raman spectra of surface carbon
and hydrocarbons on silver electrodes. Chem Phys Lett 76:448–452.
22. Swathi RS, Sebastian KL (2008) Resonance energy transfer from a dye molecule to
graphene. J Chem Phys 129:054703.
23. Xie LM, Ling X, Fang Y, Zhang J, Liu ZF (2009) Graphene as a substrate to suppress
fluorescence in resonance Raman spectroscopy. J Am Chem Soc 131:9890–9891.
24. Wang YY, et al. (2010) Gold on graphene as a substrate for surface enhanced Raman
scattering study. Appl Phys Lett 97:163111.
25. Ling X, et al. (2010) Can graphene be used as a substrate for Raman enhancement?
Nano Lett 10:553–561.
26. Ling X, Zhang J (2010) First-layer effect in graphene-enhanced Raman scattering. Small
6:2020–2025.
27. Li XS, et al. (2009) Large-area synthesis of high-quality and uniform graphene films on
copper foils. Science 324:1312–1314.
28. Jiao LY, et al. (2008) Creation of nanostructures with poly(methyl methacrylate)mediated nanotransfer printing. J Am Chem Soc 130:12612–12613.
29. Fang Y, Seong NH, Dlott DD (2008) Measurement of the distribution of site enhancements in surface-enhanced Raman scattering. Science 321:388–392.
30. Cang H, et al. (2011) Probing the electromagnetic field of a 15-nanometre hotspot by
single molecule imaging. Nature 469:385–388.
31. Cao YWC, Jin RC, Mirkin CA (2002) Nanoparticles with Raman spectroscopic fingerprints for DNA and RNA detection. Science 297:1536–1540.
32. Liu Z, et al. (2011) Revealing the molecular structure of single-molecule junctions in
different conductance states by fishing-mode tip-enhanced Raman spectroscopy.
Nat Commun 2:305.
33. Alvarez-Puebla RA, et al. (2011) Gold nanorods 3D-supercrystals as surface enhanced
Raman scattering spectroscopy substrates for the rapid detection of scrambled prions.
Proc Natl Acad Sci USA 108:8157–8161.
34. Novoselov KS, et al. (2004) Electric field effect in atomically thin carbon films. Science
306:666–669.
35. Kravets VG, et al. (2010) Spectroscopic ellipsometry of graphene and an excitonshifted van Hove peak in absorption. Phys Rev B 81–155413.
36. Schedin F, et al. (2010) Surface-enhanced Raman spectroscopy of graphene. ACS Nano
4:5617–5626.
37. Johnson PB, Christy RW (1972) Optical constants of noble metals. Phys Rev B
6:4370–4379.
Materials and Methods
9286 ∣
www.pnas.org/cgi/doi/10.1073/pnas.1205478109
Xu et al.
Download